You are on page 1of 10

Spin crossover composite materials for

electrothermomechanical actuators
Il’Ya A. Gural’Skiy, Carlos Mario Quintero, José Sánchez Costa, Philippe
Demont, Gabor Molnar, Lionel Salmon, Helena J. Shepherd, Azzedine
Bousseksou

To cite this version:


Il’Ya A. Gural’Skiy, Carlos Mario Quintero, José Sánchez Costa, Philippe Demont, Gabor Mol-
nar, et al.. Spin crossover composite materials for electrothermomechanical actuators. Journal
of Materials Chemistry C, Royal Society of Chemistry, 2014, vol. 2 (n 16), pp. 2949-2955.
<10.1039/c4tc00267a>. <hal-01166077>

HAL Id: hal-01166077


https://hal.archives-ouvertes.fr/hal-01166077
Submitted on 22 Jun 2015

HAL is a multi-disciplinary open access L’archive ouverte pluridisciplinaire HAL, est


archive for the deposit and dissemination of sci- destinée au dépôt et à la diffusion de documents
entific research documents, whether they are pub- scientifiques de niveau recherche, publiés ou non,
lished or not. The documents may come from émanant des établissements d’enseignement et de
teaching and research institutions in France or recherche français ou étrangers, des laboratoires
abroad, or from public or private research centers. publics ou privés.
Open Archive TOULOUSE Archive Ouverte (OATAO)
OATAO is an open access repository that collects the work of Toulouse researchers and
makes it freely available over the web where possible.

This is an author-deposited version published in : http://oatao.univ-toulouse.fr/


Eprints ID : 11557

To link to this article : doi: 10.1039/c4tc00267a


URL : http://dx.doi.org/10.1039/c4tc00267a

To cite this version : Gural'skiy, Il'ya A. and Quintero, Carlos M. and


Costa, José Sánchez and Demont, Philippe and Molnar, Gabor and
Salmon, Lionel and Shepherd, Helena J. and Bousseksou, Azzedine
Spin crossover composite materials for electrothermomechanical
actuators. (2014) Journal of Materials Chemistry. C, vol. 2 (n° 16). pp.
2949-2955. ISSN 2050-7526

Any correspondance concerning this service should be sent to the repository


administrator: staff-oatao@listes-diff.inp-toulouse.fr
DOI: 10.1039/c4tc00267a

Spin crossover composite materials for


electrothermomechanical actuators†
Il'ya A. Gural'skiy,ab Carlos M. Quintero,c José Sánchez Costa,a Philippe Demont,d
Gábor Molnár,a Lionel Salmon,a Helena J. Shepherd*e and Azzedine Bousseksou*a

Composites of the spin crossover complex [Fe(trz)(H-trz)2](BF4) (H-trz ¼ 1,2,4-4H-triazole and trz ¼ 1,2,4-
triazolato) dispersed in a poly(methylmethacrylate) (PMMA) matrix were synthesized and investigated for
their spin crossover properties by optical reflectivity, Raman spectroscopy and calorimetry. These
composite films were used to fabricate bilayer cantilevers that can perform efficient and tuneable
mechanical actuation based on the spin transition. A prototype device that uses the spin transition
phenomenon to convert electrical energy into mechanical motion through Joule heating is described.
This device is used to perform oscillatory actuation driven by a modulated current. The ability to tune the
performance of this electromechanical system is demonstrated by varying the working temperature, the
applied ac current and its frequency.

during the transition between their high spin (HS) and low spin
Introduction (LS) electronic congurations.7,8 Despite copious investigations
The transformation of electrical energy into motion is one of the into their magnetic and optical properties, the spontaneous
main demands from micro- and nanoelectromechanical linear strain (3) that can reach 8% (ref. 9 and 10) in these
systems (MEMS and NEMS).1 This conversion could be achieved materials seems to be one of the most fascinating assets for
via electrostatic2 or piezoelectric effects,3 Joule heating,4 etc. practical applications.11 The 3 values in SCO complexes are high
Among these methods electrothermal actuation is a widely used when compared to analogous phase transition materials. For
approach due to its simplicity both for the fabrication process example, in the very promising VO2 known for the reversible
and the actuation mechanism.5 The choice of the actuating metal–insulator phase change, a linear strain up to 2% can be
material is a crucial step in the construction of a new electro- reached.12–16 Another important asset is that SCO can be
mechanical system and the numerous specic requirements provoked by different external stimuli such as temperature,
make the development of novel actuating materials necessary.6 pressure, magnetic eld, excitation with light or adsorption of
For instance, the energy density of a mechanical actuator is chemical species. The interest is additionally supported by the
basically dependent on the strain that can be generated in its existence of hundreds of SCO complexes reported to date with
active materials. However, many other important performance different transition temperatures, hysteresis and abruptness of
indicators, such as the displacement, the force, actuation speed the spin transition.7 This versatility can be further extended by
(or frequency) and energy efficiency will also depend on the the elaboration of composite materials. For example, SCO
material selection. compounds have been imbedded in silica17–19 and various
Spin crossover (SCO) complexes of 3d4–3d7 transition metal polymer20–25 matrices providing nanoscale connement and/or
ions are well known for changing a variety of physical properties better processability.
Recently we reported on the application of spin crossover
a
LCC, CNRS & University of Toulouse (UPS, INPT), 205 route de Narbonne, 31077 materials for the elaboration of bimorph cantilevers, which
Toulouse, France. E-mail: azzedine.bousseksou@lcc-toulouse.fr were actuated both thermally and by light irradiation.11 As a
b
Department of Chemistry, Taras Shevchenko National University of Kyiv, 64 proof of concept, a single crystal of the SCO complex
Volodymyrska Street, 01601 Kyiv, Ukraine {Fe(3-CNpy)[Au(CN)2]2}$2/3H2O (py ¼ pyridine) was covered
c
LAAS, CNRS & Université de Toulouse (UPS, INSA, IAES), 7 avenue du colonel Roche, with a layer of Al and used as a cantilever actuator. A reversible
31077 Toulouse, France
d
movement of this bimorph was reached as a result of thermally
Institut Carnot CIRIMAT, CNRS & Université de Toulouse III, 118 Route de Narbonne,
31062 Toulouse, France
and photo-induced spin transitions, thus revealing a novel
e
Department of Chemistry, University of Bath, Bath, BA2 7AY, UK. E-mail: h.j. mechanism of actuation that opens up perspectives for the
shepherd@bath.ac.uk development of MEMS, articial muscles and other applica-
† Electronic supplementary information (ESI) available: Actuation of a cantilever tions for the harvesting and transformation of energy (thermal,
with a different geometry, detailed theoretical calculations, movie of the cantilever electric, photonic, .). Additionally, three bilayer polymer/SCO
oscillation. See DOI: 10.1039/c4tc00267a
composite devices, each with different SCO compounds, were and cannot be removed from the substrate or further processed.
developed to demonstrate the versatility of SCO materials for The spin transition properties of these composites were
actuating tasks. In this paper we go further by describing a observed and analyzed by optical microscopy since the SCO is
prototype device that can exploit the spin transition phenom- accompanied by a drastic change of color in the active complex
enon to convert electrical energy into motion via Joule heating from pink in the LS state to colorless in the HS state. This effect
in a composite bilayer actuator. is associated with the reversible bleaching of the characteristic
1
A1 / 1T1 absorption band centered around l ¼ 550 nm. Fig. 1
Results and discussion reveals a strong thermal dependence of the composite color that
is clearly associated with the HS 4 LS spin transition. The SCO
SCO/polymer composites in the composites takes place within the same region of
Even though the application of a single crystal SCO material for temperature as for the pure complex. The transition curves of
the construction of bimorph cantilevers was found efficient and the composites were found to be relatively independent of the
the behavior of these actuators is easy to rationalize from a polymer/complex ratio; the most signicant differences being a
theoretical point of view, the practical use of such systems has slight loss of abruptness of the spin transition for composites
certain limitations. This is mostly related to the serendipity of with lower loadings of the active material. This observation may
the crystal growth. In contrast, techniques for employing poly- simply be related to the lower thermal conductivity of the
mers and polymer composites in MEMS design are well devel- polymer component compared to the coordination compound.
oped.26 Hence, we synthesized a polymer composite material, While the SCO behaviour of composite materials is in general
consisting of the spin crossover complex [Fe(trz)(H-trz)2](BF4) not easy to predict, the negligible inuence of a polymer matrix
(form 1b, one of the two forms described by Kröber et al.27) and on those properties, as shown here, is certainly not
the organic polymer PMMA, which can produce continuous unprecedented.20–23
lms of several microns in thickness over a large area. This Selected Raman spectra of [Fe(trz)(H-trz)2](BF4) and the
iron(II) compound is known for its abrupt and robust spin PMMA/[Fe(trz)(H-trz)2](BF4) composite (50% loading) recorded
transition that takes place with a large thermal hysteresis at 20  C and 110  C are shown in Fig. 2. In the low frequency
around 85  C on heating and 60  C on cooling. It is important to region (100–300 cm1), very signicant changes are observed
note also that in this particular SCO complex the transition is upon the SCO. As discussed in ref. 29, the modes at 300, 288,
accompanied by a volume change of 11%.28 211, 202 cm1 are attributed to the LS state of the complex,
A series of PMMA/1b composites with different relative whereas those at 186 and 109 cm1 are associated with the HS
proportions of the active complex were fabricated by mechan- state. These changes of the vibrational frequencies accompa-
ical dispersion of the powder of 1b in a PMMA solution with nying the spin transition are a general trend and originate from
subsequent evaporation of the chloroform solvent. A drop- the strong electron–lattice coupling in these systems.7 The
casting procedure allows the deposition of the composite lms disappearance of the characteristic LS/HS modes at high/low
over different surfaces. Films with different proportions (10– temperatures allows us to conclude that the spin transition is
50%) of the active complex were produced (Fig. 1). Loading in virtually complete in both directions for the isolated complex as
excess of 50% of the SCO complex results in lms that are brittle well as for the composite material.

Fig. 1 Temperature dependence of the optical reflectance (l ¼ 550


nm) of [Fe(trz)(H-trz)2](BF4) (100%) and of the composite materials
PMMA/[Fe(trz)(H-trz)2](BF4) with different loadings (10–50%) of the Fig. 2 Variable-temperature Raman spectra of the [Fe(trz)(H-
active complex. Heating and cooling modes are indicated by arrows. trz)2](BF4) complex and the PMMA/[Fe(trz)(H-trz)2](BF4) (50%)
The inset shows the photographs of the composite films. composite material.
Differential scanning calorimetry (DSC) curves of [Fe(trz)(H-
trz)2](BF4) and the PMMA/[Fe(trz)(H-trz)2](BF4) composite (50%)
are shown in Fig. 3.
In agreement with the optical reectivity and Raman spec-
troscopy measurements, these data demonstrate a high degree
of similarity between the SCO in the isolated complex and the
composite materials. Shis of the transition temperatures
between the complex and its composite are attributed to ther-
malization issues – an organic polymer would be expected to be
less thermally conductive than a coordination compound. This
measurement allows also the calculation of the enthalpy (DH ¼ Fig. 4 Schematic representation of a SCO bilayer cantilever with
electrothermal actuation.
25 kJ mol1) and entropy (DS ¼ 73 J mol1 K1) changes asso-
ciated with the spin transition in 1b. These values were found
similar for the isolated complex and the complex embedded in
PMMA and they are also in good agreement with the data particles (industrially available conductive paint) to act as the
reported in ref. 27 for the pure complex. conductive layer because this conductive composite has a much
higher adhesion to the active layer than a simple metal layer can
have. Moreover, this choice for the conductive component
Bilayer actuators
makes both layers of the bimorph similar in terms of thermal
The composite lm with the highest proportion of active expansion and Young's modulus. This minimizes the mismatch
material (50%) was then used to construct bilayer actuators. The of the thermal expansion between the two layers and maximizes
highest possible loading by the active compound is expected to the deection of the strip (for a given geometry).
optimize the strain that can be transmitted from the complex to The bilayer lms were fabricated by spin-coating the
the polymer matrix, thus maximizing the actuating capabilities. conductive composite solution over a PMMA/SCO composite
Despite the proportion of the strain absorbed by the exible lm. A cantilever was subsequently cut from this bilayer and
polymer matrix, the spin transition in the active component connected to two electrodes: one electrode is xed, whereas the
results in a signicant isotropic expansion of the composite second electrode is made of a exible thin aluminium foil so as
material. to allow (nearly) free movement with the cantilever (Fig. 4).
In a simple bilayer strip, the difference in expansion of the This setup was placed in a cryostat to control the tempera-
two layers results in the deection of the cantilever that can be ture around the cantilever. The behavior of this cantilever was
exploited for actuation. In our device, the second (inert) layer is initially studied within a range of 23–100  C. Heating was
used not only to generate the strain mismatch, but is also used stopped at 100  C to avoid any transformations of the polymers
as a heating element via the Joule effect. In such a way, the (e.g. glass transition) and damage of the device. As temperature
application of a current to the cantilever should provoke a spin increases the spin transition in [Fe(trz)(H-trz)2](BF4) causes a
transition (from LS to HS), while removing the current should spontaneous strain that results in a strong deformation of the
cause a dissipation of heat followed by the reverse transition bimorph at the temperature of the spin transition. The occur-
(HS to LS). We selected a polymer that was lled with silver rence of SCO can also be clearly inferred from the simultaneous
change of the composite color during the actuation (Fig. 5). The
movement of the cantilever was quantied by following with a
video-microscope the position of the end connected to the
exible electrode. Displacement as a function of temperature is
shown in Fig. 5a.
The hysteresis loop of the thermal actuation (Fig. 5a) is
rather deformed in comparison with the spin transition loop
(Fig. 1). This is attributed to inefficient heat transfer from the
heating plate to the cantilever suspended in air, attached only at
the ends. In addition to heat exchange effects, mechanical
stress can also inuence the form of this curve. However, the
limited region of temperature as well as the hysteretic nature of
the deection clearly demonstrate that tip deection is the
result of the spin transition in the active complex.
The amplitude of actuation (ca. 1600 mm) allows us to
calculate the reactive force at the end of the cantilever and the
effective linear strain of the composite material to be 20 mN
Fig. 3 DSC curves of the pure [Fe(trz)(H-trz)2](BF4) complex and the and 1%, respectively (see ESI† for details on the calculations).
PMMA/[Fe(trz)(H-trz)2](BF4) (50%) composite material acquired at These values are slightly underestimated since the exible
heating/cooling rates of 10  C min1. electrode acts as a load on the cantilever. Nevertheless the
Fig. 6 Electric current induced actuation of a bimorph cantilever. Tip
position of the cantilever as a function of time (upper panel) and the
corresponding applied current signal (bottom panel). The frequency of
the alternative current is 0.1 Hz, Imax ¼ 300 mA, T ¼ 20  C.

additional load, it does not play a crucial role in the movement


of the cantilever. This is due to the force (20 mN) generated at
the tip of the cantilever being considerably higher than the force
Fig. 5 Thermal actuation of a bimorph cantilever. (a) Tip position of necessary to move the aluminium foil.
the cantilever as a function of temperature. Heating and cooling Aer an initial “run-in” period of a few cycles, the actuating
modes are indicated by arrows. (b) Photographs (8.6 mm  7.6 mm) of amplitude becomes highly repeatable within the given oper-
the cantilever at 68  C in the cooling and heating modes demonstrate
ating conditions (Fig. 6). We observed that this amplitude
a shape memory effect, which occurs due to the thermal hysteresis of
the spin transition. The color change associated with the spin transition depends strongly on the experimental conditions. The most
is also observed (see also ESI Movie SM1†). important contributing factor is the fraction of the complex that
undergoes SCO when the current is turned on/off. This leads to
different amplitudes of deection with different applied
effective strain is signicantly lower than that estimated from frequency, current and temperature of the surrounding
the additive effect (ca. 2%), which we relate to the partial medium. A series of experiments were designed to probe the
absorption of strain by the polymeric matrix. The efficiency of interdependence of these three parameters (Imax, f and T).
this material for actuation is described by the volumetric work The amplitude of cantilever movement is shown as a func-
density W/V, i.e. the maximal mechanical output, which can be tion of applied current in Fig. 7. A current below 200 mA does
produced by a given volume of the composite. W/V for the 50%
composite was calculated to be ca. 150 mJ cm3 (see ESI† for
details on the calculations). This value is comparable or higher
than those reported for piezo-electric, magnetostrictive and
electroactive polymer materials.30

Electrothermal actuation of the cantilever


To generate an electrothermally driven actuation, a squared
current (I) was applied to the cantilever, which engenders Joule
heating. In a typical experiment, I was changed stepwise from
0 to Imax and back to 0 at a constant frequency (f). When I ¼ Imax,
the LS / HS transition is induced, whereas removing the
current leads to the backward HS / LS transition. This
sequence of LS / HS / LS transitions provokes oscillating
motion of the cantilever (Fig. 6). The amplitude of this move-
ment is constant in numerous cycles (see also ESI Movie SM1†).
Fig. 7 Amplitude of oscillation of the bilayer cantilever as a function of
In this experiment, the device was passed through more than the applied current Imax at constant temperature and frequency (T ¼
300 actuating cycles with no observed fatigue of the actuator. It 20  C, f ¼ 0.1 Hz). Insets: tip position of the cantilever as a function of
is worth noting that though the exible electrode acts as an time for a few selected values of Imax.
Fig. 9 Amplitude of the bilayer strip oscillation as a function of
frequency at constant temperature and ac current (T ¼ 20  C, Imax ¼
300 mA). Insets: tip position of the cantilever as a function of time at
selected frequencies.

Fig. 8 Amplitude of the bilayer cantilever oscillation at different maximum value) at fcut off z 0.5 Hz (Fig. 9). In this case the
temperatures. The ac current and frequency were fixed (Imax ¼ 300 completeness of the transition in both directions (LS to HS and
mA, f ¼ 0.1 Hz). Insets: tip position of the cantilever as a function of HS to LS) is limited by the rate of heat diffusion; there is
time at selected temperatures.
insufficient time for the device to reach the required tempera-
ture to switch all active centers. For this actuator there is no
frequency limitation up to f ¼ 0.1 Hz; however, no movement is
not result in any signicant actuation of the cantilever. This observed for f > 1 Hz. This limitation is clearly related to the heat
threshold effect is likely to be the result of insufficient heating, transfer processes and not to the spin conversion process that
i.e. below a certain temperature the nucleation of the HS phase has much faster dynamics.7,8
cannot take place. The low amplitude oscillation observed for The observed cut-off frequency is not a fundamental limit, but
currents below this value should be considered chiey as a rather denotes that (if necessary) the device must be operated at a
thermal expansion effect without SCO. By contrast, a current over different temperature and/or a different current should be
200 mA generates enough heat to induce the thermal SCO, applied. Even more importantly, the cantilever geometry and the
reaching a maximum amplitude of actuation at Imax z 300 mA. thermalisation conditions can also drastically change the
The deection at currents between 200 and 300 mA corresponds performance of these actuators in terms of actuating speed. One
to a partial spin transition of the Fe(II) complex in the composite. example describing the effect of a different geometry is shown in
This threshold current value can be tuned via several different the ESI (Fig. S1–S3†). A cantilever of a smaller size was fabricated
approaches, including the change of SCO active complex. Indeed, and studied in a similar manner. A direct consequence of the size
the use of any SCO compound with higher/lower temperature of reduction is the faster heat exchange between the cantilever and
transition should increase/decrease the critical value of current its surroundings. Thus, a lower cut-off temperature and current,
necessary to drive the cantilever oscillation. as well as a higher cut-off frequency were observed.
Measurements at constant f and Imax reveal a clear dependence
of the actuation amplitude on the surrounding temperature as
shown in Fig. 8. Up to 50  C a slight and almost linear diminution Experimental information
of the actuating amplitude is observed and can be explained by an
incomplete HS to LS transition when the current is switched off. PMMA (Mw ¼ 550 000), iron(II) tetrauoroborate hexahydrate,
Above 50  C an abrupt drop off in amplitude is observed, due to 1-H-1,2,4-triazole and chloroform were purchased from Sigma
the thermal SCO occurring above this point. Thus, even when Aldrich and used without any further purication. Conductive
there is no current applied to the cantilever, a signicant fraction silver paint RS 186-3593 was purchased from RS Components.
of the active SCO complex remains in the HS state. This paint contains a polymeric basis lled with silver micro-
Again, this characteristic cut-off temperature can be tuned by spheres (<18 mm, 56%) and dissolved in an n-butyl acetate/
substituting the active complex for another SCO material. The 2-methoxy-1-methylethyl acetate mixture.
abrupt on/off switch of the actuation with temperature can be [Fe(trz)(H-trz)2](BF4) (form 1b) was prepared from
achieved by using highly cooperative spin crossover systems Fe(BF4)2$6H2O and 1-H-1,2,4-triazole according to the procedure
where the spin state change occurs sharply within < 1 K. (N.B. To described by Kröber et al.27 The active composite material was
observe an abrupt cut-off as a function of temperature efficient elaborated by mechanical dispersion of the [Fe(trz)(H-trz)2](BF4)
thermalization is also necessary.) On the other hand, the use of powder (5 mg) in a solution of PMMA (5 mg) in CHCl3 (200 ml).
compounds with a gradual SCO – taking place within hundreds This suspension was drop-casted onto a silicon substrate (1  1
degrees – can allow ne control over the actuation amplitude by cm2) and allowed to dry. 150 ml of the conductive paint was
changing the environmental temperature. subsequently spin coated (2000 rpm, 3000 rpss, 30 s) onto the
Varying the frequency at constant Imax and T reveals a cut-off upper surface. The resulting bilayer lm was removed from the
of the actuation (dened as a decay in amplitude of 0.707 of the substrate and cut into cantilevers. The cantilever used was 6.7
mm in length and 1 mm in width. The thicknesses of the SCO Notes and references
and conductive layers were ca. 100 and 50 mm, respectively.
Variable temperature Raman spectra of the samples were 1 M. Bao, in Analysis and Design Principles of MEMS Devices,
acquired by means of an Xplora (Horiba) microspectrometer Elsevier Science, Amsterdam, 2005, pp. 1–32.
and a Linkam THMS600 cryostat. A long working-distance 2 Comprehensive Microsystems, ed. Y. Gianchandani, O. Tabata
objective (50, numerical aperture ¼ 0.5) was used to focus the and H. Zappe, Elsevier, Oxford, 2008, vol. 2, pp. 1–100.
laser beam (638 nm, 0.1 mW) on the sample and to collect the 3 T.-L. Ren, H.-J. Zhao, L.-T. Liu and Z.-J. Li, Mater. Sci. Eng., B,
scattered photons. Rayleigh scattering was removed by an edge 2003, 99, 159–163.
lter. Differential scanning calorimetry analysis was carried out 4 S. Pal and H. Xie, J. Micromech. Microeng., 2012, 22, 115036.
on a Netsch DSC 204 instrument under helium purging gas. The 5 P. Srinivasan and S. M. Spearing, J. Microelectromech. Syst.,
measurements were conducted between 0 and 100  C at a scan 2008, 17, 653–667.
rate of 10 K min1 both on heating and cooling. Temperature 6 MEMS Materials and Processes Handbook, ed. R. Ghodssi and
and enthalpy were calibrated using the melting transition of P. Lin, Springer US, Boston, MA, 2011, vol. 1.
standard materials (Hg, In, Sn). 7 Spin Crossover in Transition Metal Compounds, Top. Curr.
In order to perform the electrical actuation experiments, a Chem., ed. P. Gütlich and H. A. Goodwin, Springer, Berlin,
cantilever was xed onto a xed steel electrode with the Heidelberg, 2004, vol. 233–235.
conductive paint. The other side was attached to a exible 8 A. Bousseksou, G. Molnár, L. Salmon and W. Nicolazzi,
aluminium electrode (10 mm thick) in the same manner. The Chem. Soc. Rev., 2011, 40, 3313–3335.
temperature was controlled using a Linkam HFS600E cryostat 9 P. Guionneau, Dalton Trans., 2014, 43, 382–393.
equipped with exible internal microelectrodes to connect the 10 S. Cobo, D. Ostrovskii, S. Bonhommeau, L. Vendier,
device to the current source. The videomicroscopy and optical G. Molnár, L. Salmon, K. Tanaka and A. Bousseksou, J. Am.
reectivity images were acquired with a MOTIC stereomicro- Chem. Soc., 2008, 130, 9019–9024.
scope equipped with a colour camera. Current pulses were 11 H. J. Shepherd, I. A. Gural'skiy, C. M. Quintero, S. Tricard,
generated with a programmable current source (Keithley 6430). L. Salmon, G. Molnár and A. Bousseksou, Nat. Commun.,
Tip displacement was determined from the video images. The 2013, 4, 2607.
image treatments were carried out under ImageJ and MatLab 12 R. Cabrera, E. Merced, N. Sepúlveda and F. E. Fernández, J.
environments as described previously.11 Appl. Phys., 2011, 110, 094510.
13 J. Cao, W. Fan, Q. Zhou, E. Sheu, A. Liu, C. Barrett and J. Wu,
J. Appl. Phys., 2010, 108, 083538.
Conclusions and perspectives 14 K. Liu, C. Cheng, Z. Cheng, K. Wang, R. Ramesh and J. Wu,
Nano Lett., 2012, 12, 6302–6308.
The spin crossover complex [Fe(trz)(H-trz)2](BF4) was dispersed in
15 A. Rúa, F. E. Fernández and N. Sepúlveda, J. Appl. Phys., 2010,
an organic polymer matrix to form freestanding lms. The spin
107, 074506.
crossover properties of the isolated complex were preserved in the
16 K. Wang, C. Cheng, E. Cardona, J. Guan, K. Liu and J. Wu,
composites, which allowed us the development of a highly effi-
ACS Nano, 2013, 7, 2266–2272.
cient electromechanical actuating device with a volumetric work
17 C. Faulmann, J. Chahine, I. Malfant, D. de Caro, B. Cormary
density of ca. 150 mJ cm3. The prototype device described here
and L. Valade, Dalton Trans., 2011, 40, 2480–2485.
exhibited repeatable and tunable actuation via temperature,
18 S. Titos-Padilla, J. M. Herrera, X.-W. Chen, J. J. Delgado and
frequency and current variation and provided both large force
E. Colacio, Angew. Chem., Int. Ed., 2011, 50, 3290–3293.
(several mN) and large displacement (several mm). The procedure
19 P. Durand, S. Pillet, E.-E. Bendeif, C. Carteret, M. Bouazaoui,
to elaborate the active composites can be easily extended not only
H. El Hamzaoui, B. Capoen, L. Salmon, S. Hebert,
to other SCO compounds, but also to different phase transition
J. Ghanbaja, L. Arandah and D. Schaniel, J. Mater. Chem.
materials (VO2, etc.). This approach of actuation driven by SCO can
C, 2013, 1, 1933–1942.
be also introduced in MEMS/NEMS sensors, since several SCO
20 S.-W. Lee, J.-W. Lee, S.-H. Jeong, I.-W. Park, Y.-M. Kim and
compounds have shown a good sensitivity to different hazardous
J.-I. Jin, Synth. Met., 2004, 142, 243–249.
compounds.31 An exciting perspective of this work will be the
21 A. Nakamoto, Y. Ono, N. Kojima, D. Matsumura and
investigation of the effects of scaling, since the faster heat transfer
T. Yokoyama, Chem. Lett., 2003, 32, 336–337.
rate at reduced scales will result in higher work frequency and
22 M.-L. Boillot, S. Pillet, A. Tissot, E. Rivière, N. Claiser and
effectiveness.5 This prospect is tangible since the spin transition
C. Lecomte, Inorg. Chem., 2009, 48, 4729–4736.
properties can be preserved down to very small sizes.8
23 M. Rubio, R. Hernandez, A. Nogales, A. Roig and D. Lopez,
Eur. Polym. J., 2011, 47, 52–60.
Acknowledgements 24 J. Larionova, L. Salmon, Y. Guari, A. Tokarev, K. Molvinger,
G. Molnár and A. Bousseksou, Angew. Chem., Int. Ed., 2008,
Financial support from the project Switchelec (ANR-12-IS07- 47, 8236–8240.
0003-01) is acknowledged. JSC thanks the Marie-Curie research 25 W. Hellel, A. Ould-Hamouda, J. Degert, J.-F. Létard and
program (NanoSCOpe 328078). E. Freysz, Appl. Phys. Lett., 2013, 103, 143304.
26 C. Liu, Adv. Mater., 2007, 19, 3783–3790. 29 N. O. Moussa, D. Ostrovskii, V. M. Garcia, G. Molnár,
27 J. Kröber, J.-P. Audiere, R. Claude, E. Codjovi, O. Kahn, K. Tanaka, A. B. Gaspar, J. A. Real and A. Bousseksou,
J. G. Haasnoot, F. Groliere, C. Jay and A. Bousseksou, Chem. Phys. Lett., 2009, 477, 156–159.
Chem. Mater., 1994, 6, 1404–1412. 30 V. Q. Nguyen, A. S. Ahmed and R. V. Ramanujan, Adv. Mater.,
28 A. Grosjean, P. Négrier, P. Bordet, C. Etrillard, D. Mondieig, 2012, 24, 4041–4054.
S. Pechev, E. Lebraud, J.-F. Létard and P. Guionneau, Eur. J. 31 C. Bartual-Murgui, A. Akou, L. Salmon, G. Molnár, C. Thibault,
Inorg. Chem., 2013, 796–802. J. A. Real and A. Bousseksou, Small, 2011, 7, 3385–3391.

You might also like